arxiv:math/ v1 [math.gm] 13 Dec 2005

Similar documents
Hausdorff Continuous Viscosity Solutions of Hamilton-Jacobi Equations

Walker Ray Econ 204 Problem Set 2 Suggested Solutions July 22, 2017

NAME: Mathematics 205A, Fall 2008, Final Examination. Answer Key

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

Math 209B Homework 2

Continuity. Chapter 4

MH 7500 THEOREMS. (iii) A = A; (iv) A B = A B. Theorem 5. If {A α : α Λ} is any collection of subsets of a space X, then

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

SOME REMARKS ON THE SPACE OF DIFFERENCES OF SUBLINEAR FUNCTIONS

Continuity. Chapter 4

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

EXISTENCE OF PURE EQUILIBRIA IN GAMES WITH NONATOMIC SPACE OF PLAYERS. Agnieszka Wiszniewska Matyszkiel. 1. Introduction

van Rooij, Schikhof: A Second Course on Real Functions

A Class of LULU Operators on Multi-Dimensional Arrays

THEOREMS, ETC., FOR MATH 516

MA651 Topology. Lecture 9. Compactness 2.

THEOREMS, ETC., FOR MATH 515

The homotopies of admissible multivalued mappings

Continuous Functions on Metric Spaces

7 Complete metric spaces and function spaces

4th Preparation Sheet - Solutions

An algebraic approach to Gelfand Duality

MATH 202B - Problem Set 5

z -FILTERS AND RELATED IDEALS IN C(X) Communicated by B. Davvaz

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

Chapter 2 Metric Spaces

Math General Topology Fall 2012 Homework 11 Solutions

Stanford Mathematics Department Math 205A Lecture Supplement #4 Borel Regular & Radon Measures

PROBLEMS. (b) (Polarization Identity) Show that in any inner product space

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Darboux functions. Beniamin Bogoşel

LECTURE 15: COMPLETENESS AND CONVEXITY

Math 115 Spring 11 Written Homework 10 Solutions

Fragmentability and σ-fragmentability

The ideal of Sierpiński-Zygmund sets on the plane

INTRODUCTION TO TOPOLOGY, MATH 141, PRACTICE PROBLEMS

Economics 204 Summer/Fall 2011 Lecture 5 Friday July 29, 2011

Notes for Functional Analysis

3 Measurable Functions

CHRISTENSEN ZERO SETS AND MEASURABLE CONVEX FUNCTIONS1 PAL FISCHER AND ZBIGNIEW SLODKOWSKI

Formal Groups. Niki Myrto Mavraki

SOME QUESTIONS FOR MATH 766, SPRING Question 1. Let C([0, 1]) be the set of all continuous functions on [0, 1] endowed with the norm

Bounded uniformly continuous functions

Math 426 Homework 4 Due 3 November 2017

Applied Analysis (APPM 5440): Final exam 1:30pm 4:00pm, Dec. 14, Closed books.

Measurable Choice Functions

B. Appendix B. Topological vector spaces

Bounded and continuous functions on a locally compact Hausdorff space and dual spaces

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond

n [ F (b j ) F (a j ) ], n j=1(a j, b j ] E (4.1)

MTG 5316/4302 FALL 2018 REVIEW FINAL

Aliprantis, Border: Infinite-dimensional Analysis A Hitchhiker s Guide

Multivector Functions of a Multivector Variable

REVIEW OF ESSENTIAL MATH 346 TOPICS

Analysis Qualifying Exam

P.S. Gevorgyan and S.D. Iliadis. 1. Introduction

Topological vectorspaces

4 Countability axioms

ANALYSIS WORKSHEET II: METRIC SPACES

THIRD SEMESTER M. Sc. DEGREE (MATHEMATICS) EXAMINATION (CUSS PG 2010) MODEL QUESTION PAPER MT3C11: COMPLEX ANALYSIS

Combinatorics in Banach space theory Lecture 12

Topological Properties of Operations on Spaces of Continuous Functions and Integrable Functions

ON FUNCTIONS WHOSE GRAPH IS A HAMEL BASIS

REAL AND COMPLEX ANALYSIS

Max-min (σ-)additive representation of monotone measures

Test 2 Review Math 1111 College Algebra

LECTURE 6. CONTINUOUS FUNCTIONS AND BASIC TOPOLOGICAL NOTIONS

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

A NOTE ON MULTIPLIERS OF SUBTRACTION ALGEBRAS

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

Some Background Material

Limit Theorems. MATH 464/506, Real Analysis. J. Robert Buchanan. Summer Department of Mathematics. J. Robert Buchanan Limit Theorems

Mathematical Methods for Physics and Engineering

(1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define

NORMAL FAMILIES OF HOLOMORPHIC FUNCTIONS ON INFINITE DIMENSIONAL SPACES

The Hopf argument. Yves Coudene. IRMAR, Université Rennes 1, campus beaulieu, bat Rennes cedex, France

h(x) lim H(x) = lim Since h is nondecreasing then h(x) 0 for all x, and if h is discontinuous at a point x then H(x) > 0. Denote

Generalized directional derivatives for locally Lipschitz functions which satisfy Leibniz rule

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Real Analysis Problems

CHAPTER V DUAL SPACES

Math 321 Final Examination April 1995 Notation used in this exam: N. (1) S N (f,x) = f(t)e int dt e inx.

PREVALENCE OF SOME KNOWN TYPICAL PROPERTIES. 1. Introduction

Measurability of Intersections of Measurable Multifunctions

Chapter 5. Measurable Functions

IDEAL CONVERGENCE OF DOUBLE INTERVAL VALUED NUMBERS DEFINED BY ORLICZ FUNCTION

Three hours THE UNIVERSITY OF MANCHESTER. 24th January

Chapter 4. Measurable Functions. 4.1 Measurable Functions

FUNCTIONAL ANALYSIS HAHN-BANACH THEOREM. F (m 2 ) + α m 2 + x 0

THE NEARLY ADDITIVE MAPS

A C 0 coarse structure for families of pseudometrics and the Higson-Roe functor

arxiv: v1 [math.fa] 14 Jul 2018

Nel s category theory based differential and integral Calculus, or Did Newton know category theory?

MATH 54 - TOPOLOGY SUMMER 2015 FINAL EXAMINATION. Problem 1

Most Continuous Functions are Nowhere Differentiable

REAL VARIABLES: PROBLEM SET 1. = x limsup E k

Iowa State University. Instructor: Alex Roitershtein Summer Homework #5. Solutions

1.4 The Jacobian of a map

CHAPTER I THE RIESZ REPRESENTATION THEOREM

FIRST ASSIGNMENT. (1) Let E X X be an equivalence relation on a set X. Construct the set of equivalence classes as colimit in the category Sets.

Transcription:

arxiv:math/0512272v1 [math.gm] 13 Dec 2005 Algebraic operations on the space of Hausdorff continuous interval functions Roumen Anguelov Department of Mathematics and Applied Mathematics University of Pretoria, SOUTH AFRICA roumen.anguelov@up.ac.za Blagovest Sendov Bulgarian Embassy, 36-3,Yoyogi 5-chome, Shibuia-ku, Tokyo 151-0053 JAPAN sendov2003@yahoo.com Svetoslav Markov Institute of Mathematics and Informatics, BAS Acad. G. Bonchev st., block 8, 1113 Sofia, BULGARIA smarkov@bio.bas.bg Abstract We show that the operations addition and multiplication on the set C(Ω) of all real continuous functions on Ω R n can be extended to the set H(Ω) of all Hausdorff continuous interval functions on Ω in such a way that the algebraic structure of C(Ω) is preserved, namely, H(Ω) is a commutative ring with identity. The operations on H(Ω) are defined in three different but equivalent ways. This allow us to look at these operations from different points of view as well as to show that they are naturally associated with the Hausdorff continuous functions. 1 Introduction The set H(Ω) of all Hausdorff continuous interval functions appears naturally in the context of Hausdorff approximations of real functions. The concept of Hausdorff continuity of interval functions generalizes the concept of continuity of real functio n in such a way that many essential properties of the usual continuous real functions are preserved. Not least this is due to the fact that the Hausdorff continuous functions assume real (point) values on a dense subset of the domain and are completely determined by these values. It is well known that the set C(Ω) of all continuous real functions defined on a subset Ω of R n is a commutative ring with respect to the point-wise defined addition and 1

multiplication of functions. Hence the natural question: Is it possible to extend the algebraic operations on C(Ω) to the set H(Ω) of all Hausdorff continuous functions defined on Ω in a way which preserves the algebraic structure, that is, the set of H(Ω) is a commutative ring with respect to the extended operations? We show in this paper that the answer to this question is positive. Furthermore, we give three different but equivalent ways of defining algebraic operations on H(Ω) with the required properties. Namely, (i) through the point-wise interval operations, (ii) by using an extension property of the Hausdorff continuous functions, (iii) through the order convergence structure on H(Ω). 2 General Setting The real line is denoted by R and the set of all finite real intervals by IR = {[a, a] : a, a R, a a}. Given an interval a = [a, a] = {x : a x a} IR, w(a) = a a is the width of a, while a = max{ a, a } is the modulus of a. An interval a is called proper interval, if w(a) > 0 and point interval, if w(a) = 0. Identifying a R with the point interval [a, a] IR, we consider R as a subset of IR. We denote by A(Ω) the set of all locally bounded interval-valued functions defined on an arbitrary set Ω R n. The set A(Ω) contains the set A(Ω) of all locally bounded real functions defined on Ω. Recall that a real function or an interval-valued function f defined on Ω is called locally bounded if for every x Ω there exist δ > 0 and M R such that f(y) < M, y B δ (x), where B δ (x) = {y Ω : x y < δ} denotes the open δ-neighborhood of x in Ω. Let D be a dense subset of Ω. The mappings I(D, Ω, ), S(D, Ω, ) : A(D) A(Ω) defined for f A(D) and x Ω by I(D, Ω, f)(x) = sup inf{f(y) : y B δ (x) D}, δ>0 S(D, Ω, f)(x) = inf δ(x) D}, δ>0 are called lower and upper Baire operators, respectively. The mapping F : A(D) A(Ω), called graph completion operator, is defined by F(D, Ω, f)(x) = [I(D, Ω, f)(x), S(D, Ω, f)(x)], x Ω, f A(D). In the case when D = Ω the sets D and Ω will be omitted, thus we write I(f) = I(Ω, Ω, f), S(f) = S(Ω, Ω, f), F(f) = F(Ω, Ω, f). Definition 1 A function f A(Ω) is S-continuous, if F(f) = f. Definition 2 A function f A(Ω) is Hausdorff continuous (H-continuous), if g A(Ω) with g(x) f(x), x Ω, implies F(g)(x) = f(x), x Ω. Theorem 1 [1], [10] For every f A(Ω) the functions F(I(S(f))) and F(S(I(f))) are H-continuous. The following theorem states an essential property of the continuous functions which is preserved by the H-continuity, [1]. 2

Theorem 2 Let f, g H(Ω). If there exists a dense subset D of Ω such that f(x) = g(x), x D, then f(x) = g(x), x ω. H-continuous functions are also similar to the usual continuous real functions in that they assume point values everywhere on Ω except for a set of first Baire category. More precisely, it is shown in [1] that for every f H(Ω) the set W f = {x Ω : w(f(x)) > 0} (1) is of first Baire category and f is continuous on Ω \ W f. Since a finite or countable union of sets of first Baire category is also a set of first Baire category we have: Theorem 3 Let the set Ω be open and let F be a finite or countable set of H-continuous functions. Then the set D F = {x Ω : w(f(x)) = 0, f F} = Ω \ W f (2) f F is dense in Ω and all functions f F are continuous on D F. The graph completion operator is inclusion isotone i) w. r. t. the functional argument, that is, if f, g A(D), where D is dense in Ω, then f(x) g(x), x D = F(D, Ω, f)(x) F(D, Ω, g)(x), x Ω, (3) and, ii) w. r. t. the set D in the sense that if D 1 and D 2 are dense subsets of Ω and f A(D 1 D 2 ) then D 1 D 2 = F(D 1, Ω, f)(x) F(D 2, Ω, f)(x), x Ω. (4) In particular, (4) implies that for any dense subset D of Ω and f A(Ω) we have F(D, Ω, f)(x) F(f)(x), x Ω. (5) Let f A(Ω). For every x Ω the value of f is an interval [f(x), f(x)] IR. Hence, f can be written in the form f = [f, f] where f, f A(Ω) and f(x) f(x), x Ω. The lower and upper Baire operators as well as the graph completion operator of an interval-valued function f can be represented in terms of f and f, namely, for every dense subset D of Ω: I(D, Ω, f) = I(D, Ω, f), S(D, Ω, f) = S(D, Ω, f), F(D, Ω, f) = [I(D, Ω, f), S(D, Ω, f)]. 3 Interval operations and operations on H(Ω) We recall that the operations of addition and multiplication on the set of real intervals IR are defined in more then one way, [?]. Here we consider the so called outer operations which are inclusion isotone. For [a, a], [b, b] I R [a, a] + [b, b]={a + b : a [a, a], b [b, b]}=[a + b, a + b] [a, a] [b, b]={ab:a [a, a], b [b, b]}=[min{ab, ab, ab, ab}, max{ab, ab, ab, ab}] 3

The operations for interval functions are defined point-wise in the usual way: (f + g)(x) = f(x) + g(x), (f g)(x) = f(x) g(x) (6) Example 1 Consider the functions f, g H(R) given by 0, if x < 0, 0, if x < 0, f(x) = [0, 1], if x = 0, g(x) = [ 1, 0], if x = 0, 1, if x > 0; 1, if x > 0. Using (6) we have (f + g)(x) = 0, if x < 0, [ 1, 1], if x = 0, 0, if x > 0. The considered example shows that the point-wise sum of H-continuous functions is not necessarily an H-continuous function. Hence the significance of the following two theorems. Theorem 4 [3] If the interval functions f, g A(Ω) are S-continuous then the function f + g and f g are also S-continuous. Theorem 5 Let f, g H(Ω). (a) There exists a unique function p p(x) (f + g)(x), x Ω. (b) There exists a unique function q q(x) (f g)(x), x Ω. H(Ω) satisfying the inclusion H(Ω) satisfying the inclusion Proof. We will prove only (a) because (b) is proved in a similar way. The existence of the function p follows from Theorem 1. Indeed, both functions F(I(S(f + g))) and F(S(I(f + g))) satisfy the required inclusion. To prove the uniqueness let us assume that p 1, p 2 H(Ω) both satisfy the inclusion in (a). Consider the set D fg = {x Ω : w(f(x)) = w(g(x)) = 0}. Obviously w((f + g)(x)) = 0 for x D fg. Therefore, due to the assumed inclusions we have p 1 (x) = p 2 (x) = f(x) + g(x), x D fg. According to Theorem 1 the set D fg is dense in Ω. Using that the functions p 1 and p 2 are H-continuous it follows from Theorem 2 that p 1 = p 2. Now we define the operations and as follows. Definition 3 Let f, g H)(Ω). Then f g is the unique Hausdorff continuous function satisfying the inclusion (f g)(x) (f+g)(x), x Ω; f g is the unique Hausdorff continuous function satisfying the inclusion (f g)(x) (f g)(x), x Ω. 4

The existence of both f g and f g is guaranteed by Theorem 5. It is important to note that the values of f g and f g at the points where both operands assume interval values can not be determined point-wise, i.e., from the values of f and g at these points. This is illustrated by the following example. Example 2 Consider the functions f, g H(R) given by f(x) = g(x) = { sin(1/x), if x 0, [ 1, 1], if x = 0; { cos(1/x), if x 0, [ 1, 1], if x = 0. We have (f g)(x) = { 2cos(1/x + π/4), if x 0, [ 2, 2], if x = 0. Clearly (f g)(0) can not be obtained just from the values f(0) and g(0). According to Theorem 1 for any f, g H)(Ω) the functions F(S(I(f + g))) and F(I(S(f + g))) are Hausdorff continuous. Moreover, these functions satisfy the inclusions F(S(I(f + g)))(x) (f + g)(x), F(I(S(f + g))(x) (f + g)(x), x Ω. Therefore they both coincide with f g. Hence we have f g = F(I(S(f+g))) = F(I(S(f +g)). In a similar way we obtain f g = F(I(S(f g))) = F(I(S(f g)). One can immediately see from the above representations that if f and g are usual continuous real functions we have f g = f + g and f g = f g. Hence and extend the operations of addition and multiplication on C(Ω). A further motivation for considering these operations is in the fact that the algebraic structure of C(Ω) is preserved as stated in the next theorem. Theorem 6 The set H(Ω) is a commutative ring with identity with respect to the operations and. Proof. The commutative laws for both and follow immediately from Definition 3. It is also obvious that the additive identity is the constant zero function while the multiplicative identity is the constant function equal to 1. We will show now the existence of the additive inverse. Let f = [f, f] H(Ω). Consider the function g H(Ω) given by g(x) = [ f(x), f(x)], x Ω. Clearly we have 0 f(x) + g(x), x Ω. Then according to Definition 3 f g is the constant zero function. The proof of the associative and distributive laws is an easy application of the techniques derived in the next section and will be proved there. 5

4 Extension property and an alternative definition of the operations on H(Ω) Let D be a dense subset of Ω. Extending a function f defined on D to Ω while preserving its properties (e.g. linearity, continuity) is an important issue in functional analysis. Recall that if f is continuous on D it does not necessarily have a continuous extension on Ω. Hence the significance of the next theorem which was proved in [4]. Theorem 7 Let ϕ H(D), where D is a dense subset of Ω. Then there exists a unique f H(Ω), such that f(x) = ϕ(x), x D. Namely, f = F(D, Ω, ϕ). For every two functions f, g H(Ω) denote D fg = {x Ω : w(f(x)) = w(g(x)) = 0}. As shown already the point-wise sum and product of H-continuous functions is not always H-continuous. However, the restrictions of f, g, f+g and f g on the set D fg are all real continuous functions, see Theorem 3. As usual these restrictions are denoted by f Dfg, (f +g) Dfg, etc. Using that the set D fg is dense in Ω the following definition of the operations and is suggested. Definition 4 Let f, g H(Ω). Then (a) f g is the unique H-continuous extension of (f + g) Dfg on Ω given by Theorem 7, that is, f g = F(D fg, Ω, f + g). (b) f g is the unique H-continuous extension of (f g) Dfg on Ω given by Theorem 7, that is, f g = F(D fg, Ω, f g). In order to justify the use of the notations and let us immediately prove that Definitions 3 and 4 are equivalent. Indeed, using the property (5) and the fact that f + g is S-continuous, see Theorem 4, we have F(D fg, Ω, f + g)(x) F(f + g)(x) = (f + g)(x), x Ω. Hence F(D fg, Ω, f + g) is the unique Hausdorff continuous function satisfying the inclusion required in Definition 3 which implies that F(D fg, Ω, f + g) is the sum of f and g according to Definition 3. In a similar way one can show that F(D fg, Ω, f g) is the product of f and g according to Definition 3 Therefore Definitions 3 and 4 are equivalent. Definition 4 is particularly useful for evaluating arithmetical expressions involving more than two operands since one can evaluate the expression on a set where all operands assume point values and then extend the answer to Ω. We will explain the procedure in detail. Let E(+,, z 1, z 2,..., z k ) be an expression involving the operations + and and k operands. Let F = {f 1, f 2,..., f k } H(Ω). Then the functions E(+,, f 1, f 2,..., f k ) and E(,, f 1, f 2,..., f k ) (7) are both well defined, the first one being S-continuous, the second one being H-continuous. According to Theorem 3 the set D F given by (2) is dense in 6

Ω. Then a simple connection between the functions (7) is given in the next theorem, which is proved easily using the definition of the operations and the extension property. Theorem 8 For any set of functions F = {f 1, f 2,..., f k } H(Ω) and arithmetical expression E(+,, z 1, z 2,..., z k ) we have F(D F, Ω, E(+,, f 1, f 2,..., f k )) = E(,, f 1, f 2,..., f k ) As an application of Theorem 8 we will show that the associative and distributive laws of a ring hold true on H(Ω) with respect to the operations and. This will complete the proof of Theorem 6. Proof of the associative and distributive laws on H(Ω). Let f, g, h H(Ω) and let D F be the dense subset of Ω given by (2) with F = {f, g, h}. Since the values of f, g and h on the set D F are all real numbers (point intervals) we have the functions f, g and h satisfy the associative and distributive laws with respect to the operations + and. Then using Theorem 8 we have (f g) h = F(D F, Ω, (f + g) + h) = F(D F, Ω, f + (g + h)) = f (g h), (f g) h = F(D F, Ω, (f g) h) = F(D F, Ω, f (g h)) = f (g h), (f(x) g(x)) h(x) = F(D F, Ω, (f + g) h) = F(D F, Ω, f h + g h) = f h + g h, which shows that both associative laws as well as the distributive law hold true. 5 Definition of the operations on H(Ω) through the order convergence structure Partial order can be defined for intervals in different ways. Here we consider the partial order on IR given by [a, a] [b, b] a b, a b. (8) The partial order on H(Ω) which is induced by (8) in a point-wise way, that is, for f, g H(Ω) f g f(x) g(x), x Ω, (9) is naturally associated with H(Ω). For example, it was shown in [1] that the set H(Ω) is Dedekind order complete with respect to this order. The order convergence of sequences on a poset is defined through the partial order. Definition 5 Let P be a poset with a partial order. A sequence (f n ) n N on P is said to order converge to f P if there exist on P an increasing sequence (α n ) n N and a decreasing sequence (β n ) n N such that α n ξ n β n, n N, and f = sup n N α n = inf n N β n. 7

It is well known that in general the order convergence on a poset is not topological in the sense there there is no topology with class of convergent sequences exactly equal to the class of the order convergent sequences. In particular this is the case of H(Ω) with respect to the partial order (9). However, the order convergence induces on H(Ω) the structure of the so called FS sequential convergence space. See [7] for the the definition and further details on FS sequential convergence spaces and convergence (filter) spaces. The concept of Cauchy sequence in general can not be defined within the realm of sequential convergence only but rather using the stronger concept of a convergence space, [7]. It was shown in [5] that the order convergence structure on C(Ω)) is induced by a convergence space and we have the following characterization of the Cauchy sequences. Let (f n ) n N be a sequence on C(Ω). Then (f n ) n N is Cauchy There exists a decreasing sequence (β n ) n N on C(Ω) such that inf β n = 0 and f m f k β n, m, k n, n N. (10) It was also shown in [5] that the convergence space completion of C(Ω) is H(Ω). More precisely, we have the following theorem. Theorem 9 (i) For every Cauchy sequence (f n ) n N on C(Ω) there exists f H(Ω) such that f n f. (ii) For every f H(Ω) there exists a Cauchy sequence (f n ) n N on C(Ω) such that f n f. Moreover, the sequence (f n ) n N can be selected to be either increasing or decreasing. Definition 6 Let f, g H(Ω) and let (f n ) n N and (g n ) n N be the Cauchy sequences on C(Ω) existing in terms of Theorem 9, that is, we have f n f, g n g. Then f g is the order limit of (f n + g n ) n N and f g is the order limit of (f n g n ) n N. Let us first note that the order limits stated in Definition 6 do exist. Indeed, since (f n ) n N and (g n ) n N are Cauchy sequences on C(Ω) one can see from (10) that their sum (f n + g n ) n N and their product (f n g n ) n N are Cauchy sequences. Hence according to Theorem 9(i) they both order converge. To establish the consistency of the Definition 6 we need to show that f g and f g do not depend on the particular choice of the sequences. Let (f n (1) ) n N, (f n (2) ) n N, (g n (1) ) n N, (g n (2) ) n N be Cauchy sequences on C(Ω) such that f n (i) f, g n (i) g, i = 1, 2. We will show that the sequences (f n (1) + g n (1) ) n N and (f n (2) + g n (2) ) n N converge to the same limit and that the sequences (f n (1) g n (1) ) n N and (f n (2) g n (2) ) n N converge to the same limit. Denote by (f n ) n N and (g n ) n N the trivial mixtures of the sequences (f n (1) ) n N, (f n (2) ) n N and, respectively, (g n (1) ) n N, (g n (2) ) n N, that is, f 2n 1 = f n (1), f 2n = f n (2), g 2n 1 = g n (1), g 2n = g n (2). In an FS sequential convergence space the trivial mixture of 8

sequences converging to the same limit also converges to that limit, [7]. Hence we have f n f, g n g. It easy to see that the sequences (f n ) n N, (g n ) n N are Cauchy. Hence the sequence (f n + g n ) n N is also Cauchy, which implies that (f n + g n ) n N converges on H(Ω), see Theorem 9(i). Therefore, (f n (1) + g n (1) ) n N n ) n N, being subsequences of the order convergent sequence (f n + and (f n (2) +g (2) g n ) n N order converge to the same limit. In a similar way we prove that (f (1) g (1) n ) n N and (f (2) n g (2) n ) n N converge to the same limit. Theorem 10 Definition 3 and Definition 6 are equivalent. n Proof. Let f, g H(Ω) and let h 1 be their sum according to Definition 3 while h 2 is their sum according to Definition 6. We will show that h 1 = h 2. It follows from Theorem 9(ii) that we can select increasing sequences (f n ) n N, (g n ) n N on C(Ω) such that f n f, g n g or equivalently, f = sup f n, g = sup g n. n N n N Then, according to Definition 6, h 2 is the order limit of the increasing sequence (f n + g n ) n N, that is, h 2 = sup n N (f n + g n ). On the other hand f n (x) + g n (x) f(x) + g(x) h 1 (x), x Ω. Therefore, h 2 h 1. In similar way by using decreasing sequences we prove that h 2 h 1. Hence h 1 = h 2. The proof of the equivalence of definitions of multiplication is done using a similar approach but is technically more complicated an will be omitted. References [1] Anguelov, R.: Dedekind order completion of C(X) by Hausdorff continuous functions. Quaestiones Mathematicae 27(2004) 153 170. [2] Anguelov, R.: An Introduction to some Spaces of Interval Functions, Technical Report UPWT2004/3, University of Pretoria, 2004. [3] Anguelov, R., Markov, S.: Extended Segment Analysis. Freiburger Intervall- Berichte 10 (1981) 1 63. [4] Anguelov, R., Markov, S., Sendov, B.: On the Normed Linear Space of Hausdorff Continuous Functions, Lecture Notes in Computer Science, to appear. [5] Anguelov, R., van der Walt, J. H.: Order Convergence Structure on C(X), Quaestiones Mathematicae, to appear. [6] R. Baire, Lecons sur les Fonctions Discontinues, Collection Borel, Paris, 1905. [7] R. Beattie and H.-P. Butzmann, Convergence structures and applications to functional analysis, Kluwer Academic Plublishers, Dordrecht, Boston, London, 2002. 9

[8] Markov, S.: On Quasilinear Spaces of Convex Bodies and Intervals, J. Comput. Appl. Math. 162 (2004), 93 112. [9] Rolewicz, S.: Metric Linear Spaces, PWN Polish Scientific Publishers, Warsaw, D. Reidel Publ., Dordrecht, 1984. [10] Sendov, Bl.: Hausdorff Approximations, Kluwer, 1990. 10